Diastereoselective Formation of Ruthenium-Isocyanide and -Acetylide

Apr 23, 2005 - The reaction of ruthenium complexes [(η5:η1-2-Me-4-R-C5H2CO2CH2CH2PPh2)Ru(PR'3)(NCMe)][PF6] (R = Me, Ph, But; R' = Ph, Me, OPh) posse...
7 downloads 8 Views 220KB Size
Organometallics 2005, 24, 2747-2754

2747

Diastereoselective Formation of Ruthenium-Isocyanide and -Acetylide Complexes with Planar-Chiral Cyclopentadienyl-Phosphine Ligands Yuji Matsushima, Kiyotaka Onitsuka,* and Shigetoshi Takahashi The Institute of Scientific and Industrial Research, Osaka University, 8-1 Mihogaoka, Ibaraki, Osaka 567-0047, Japan Received January 25, 2005

The reaction of ruthenium complexes [(η5:η1-2-Me-4-R-C5H2CO2CH2CH2PPh2)Ru(PR′3)(NCMe)][PF6] (R ) Me, Ph, But; R′ ) Ph, Me, OPh) possessing planar-chiral cyclopentadienylphosphine ligands with tert-butyl isocyanide resulted in the diastereoselective formation of ruthenium-isocyanide complexes (up to >99% de) under kinetic control. The reaction with phenylacetylene followed by treatment with alumina gave acetylide complexes diastereoselectively (up to >99% de). The configuration of the products was determined by X-ray analyses as well as NOE measurements. In both reactions, the stereochemistry of the ruthenium atom in the substrate did not influence the diastereoselectivity of the products, although the substituent on the cyclopentadienyl ring and the phosphorus ligands on the ruthenium atom showed steric effects on the selectivity. Introduction The synthesis and reaction of chiral organometallic complexes have attracted much attention because they provide useful information for the development of new asymmetric catalysts and for furthering our understanding of the mechanism of catalytic asymmetric reactions.1 Although organometallic complexes with chiral ligands comprise the majority of such chemistry, other types of chiral complexes have become the focus of increasing interest in recent years. A half-sandwich complex with a three-legged piano stool structure generates metal-centered chirality when the ligands differ from each other.2 Although a significant number of metal-centered chiral complexes have been prepared so far, the control of the stereochemistry at the metal center remains a challenging task.3 We have demonstrated that planar-chiral cyclopentadienyl-ruthenium (Cp′Ru) complexes with a tethered (1) Comprehensive Asymmetric Catalysis I-III; Jacobsen, E. N., Pflatz, A., Yamamoto, H., Eds.; Springer-Verlag: Berlin, 1999. (2) For recent reviews, see: (a) Brunner, H. Angew. Chem., Int. Ed. 1999, 38, 1195. (b) Brunner, H. Eur. J. Inorg. Chem. 2001, 905. (c) Ganter, C. Chem. Soc. Rev. 2003, 32, 130. (3) For recent references, see: (a) Drommi, D.; Faraone, F.; Francio, G.; Belletti, D.; Graiff, C.; Tiripicchio, A. Organometallics 2002, 21, 761. (b) Murata, K.; Konishi, H.; Ito, M.; Ikariya, T. Organometallics 2002, 21, 253. (c) Ritleng, V.; Pfeffer, M.; Sirlin, C. Organometallics 2003, 22, 347. (d) Faller, J. W.; D’Alliessi, D. G. Organometallics 2003, 22, 2749. (e) Brunner, H.; Zwack, T.; Zabel, M.; Beck, W.; Bo¨hm, A. Organometallics 2003, 22, 1741. (f) Pinto, P.; Marconi, G.; Heinemann, F. W.; Zenneck, U. Organometallics 2004, 23, 374. (g) Davenport, A. J.; Davies, D. L.; Fawcett, J.; Russell, D. R. Dalton Trans. 2004, 1481. (h) Kataoka, Y.; Nakagawa, Y.; Shibahara, A.; Yamagata, T.; Mashima, K.; Tani, K. Organometallics 2004, 23, 2095. (i) Doppiu, A.; Englert, U.; Salzer, A. Chem. Commun. 2004, 2166. (j) Alezra, V.; Bernardinelli, G.; Corminboeuf, C.; Frey, U.; Ku¨ndig, E. P.; Merbach, A. E.; Saudan, C. M.; Viton, F.; Weber, J. J. Am. Chem. Soc. 2004, 126, 4843. (k) Kataoka, Y.; Shimada, K.; Goi, T.; Yamagata, T.; Mashima, K.; Tani, K. Inorg. Chim. Acta 2004, 357, 2965. (l) Brunner, H.; Ko¨llnberger, A.; Mehmood, A.; Tsuno, T.; Zabel, M. Organometallics 2004, 23, 4006. (m) Cayuela, E.; Jalo´n, F. A.; Manzano, B. R.; Espino, G.; Weissensteiner, W.; Mereiter, K. J. Am. Chem. Soc. 2004, 126, 7049.

phosphine ligand have the ability to control stereochemistry in stoichiometric and catalytic reactions.4 For example, the reactions of the planar-chiral Cp′Ru complexes with phosphine resulted in the highly selective induction of the metal-centered chirality in the resulting complexes. Because a catalytic reaction involves multistep stoichiometric reactions, further reactions of the metal-centered chiral complexes are of special interest. From this viewpoint, we have recently reported that the reactions of the planar-chiral Cp′Ru complexes with an iodide ligand gave iodo complexes, the metal-centered chirality of which was thermodynamically controlled.5 We present herein the diastereoselective formation of ruthenium-isocyanide complexes by the ligand exchange reaction of the planar-chiral Cp′Ru complexes under kinetic control. We also examined the reaction with phenylacetylene to give vinylidene complexes, which were isolated as acetylide complexes by treatment with alumina with high diastereoselectivity. Although we used a racemic mixture for each diastereomer of the starting complexes, all structures are given with planar chirality with SCp for clarity in this article. Results Treatment of the diastereomerically pure ruthenium complex [(η5:η1-2,4-Me2-C5H2CO2CH2CH2PPh2)Ru(PPh3)(4) (a) Matsushima, Y.; Onitsuka, K.; Takahashi, S. Chem. Lett. 2000, 760. (b) Onitsuka, K.; Dodo, N.; Matsushima, Y.; Takahashi, S. Chem. Commun. 2001, 521. (c) Onitsuka, K.; Ajioka, Y.; Matsushima, Y.; Takahashi, S. Organometallics 2001, 20, 3274. (d) Matsushima, Y.; Onitsuka, K.; Kondo, T.; Mitsudo, T.; Takahashi, S. J. Am. Chem. Soc. 2001, 123, 10405. (e) Matsushima, Y.; Onitsuka, K.; Takahashi, S. Organometallics 2004, 23, 2439. (f) Matsushima, Y.; Onitsuka, K.; Takahashi, S. Organometallics 2004, 23, 3763. (5) Matsushima, Y.; Onitsuka, K.; Takahashi, S. Dalton Trans. 2004, 547.

10.1021/om0500523 CCC: $30.25 © 2005 American Chemical Society Publication on Web 04/23/2005

2748

Organometallics, Vol. 24, No. 11, 2005

Matsushima et al. Table 1. Reactions of 1-I and/or 1-II with tert-Butyl Isocyanidea

Scheme 1

(NCMe)][PF6] (1a-I), the configuration of which is SCpSRu/RCpRRu,6 with 10 equiv of tert-butyl isocyanide in refluxing dichloromethane for 12 h led to the ligand exchange reaction to give isocyanide complex [(η5:η1-2, 4-Me2-C5H2CO2CH2CH2PPh2)Ru(PPh3)(CNBut)][PF6] (2a) in 85% yield (Scheme 1). The 31P NMR spectrum of 2a clearly showed that the product consisted of two diastereomers, 2a-I (SCpSRu/RCpRRu) and 2a-II (SCpRRu/ RCpSRu), with 44% de,6 and the configuration of the major product was determined by differential NOE measurement. The NOE signals of two protons on the Cp′ ring were observed by irradiation of the signal assigned to the methyl group at the 4-position of the Cp′ ring, whereas irradiation of the signal assigned to the methyl group at the 2-position of the Cp′ ring produced NOE signals not only of the Cp′ proton at the 3-position but also of the But protons on isocyanide. These results clearly suggest that the isocyanide is situated close to the methyl group at the 2-position of the Cp′ ring in the major product. Thus, the major isomer is 2a-II and the minor one is 2a-I. When a mixture of the two diastereomers 1a-I and 1a-II (75:25) was used as the starting material, complexes 2a-I and 2a-II in a 27:73 ratio (46% de) were produced in 83% yield, suggesting that the stereochemistry of the starting material does not influence the diastereoselectivity of the product. This is similar to our previous results in the reaction of complex 1 with Bu4NI,5 but is in sharp contrast to the analogous reactions of ruthenium complexes having chiral groups on the Cp′ ring or on the phosphine ligand, which proceed with retention of configuration at the metal center.7 Then, we examined the reactions of several ruthenium complexes, the representative results of which are given in Table 1, along with the results described above. The reaction of complex 1b-I, which has a phenyl group at the 4-position of the Cp′ ring, produced isocyanide complexes 2b-I and 2b-II with 80% de (entry 3), and the reaction of a mixture of 1b-I and 1b-II in a 90:10 ratio gave essentially the same result (entry 4). By (6) The assignment for the absolute configuration at the metal center (R or S) in complexes with trimethylphosphine is opposite those of triphenylphosphine and triphenylphosphite with the same structure due to the difference in the sequence of the tethered phosphine and the phosphorus ligand. (7) (a) Morandini, F.; Consiglio, G.; Lucchini, V. Organometallics 1985, 4, 1202. (b) Cesarotti, E.; Walker, M. A. P. C.; Hursthouse, M. B.; Vefghi, R.; Schofield, P. A.; White, C. J. Organomet. Chem. 1985, 286, 343.

entry

substrate

product

yield/%

de/%b,c

1 2 3 4 5 6 7 8 9 10 11 12 13

1a-I 1a-I + 1a-II (75:25) 1b-I 1b-I + 1b-II (90:10) 1c-I 1d-I 1d-I + 1d-II (50:50) 1d-II 1e-I 1f-I 1g-I 1h-I 1i-I

2a 2a 2b 2b 2c 2d 2d 2d 2e 2f 2g 2h 2i

85 83 90 86 87 93 90 88 88 89 85 85 81

44 (II) 46 (II) 80 (II) 80 (II) >99 (I) 44 (II) 46 (II) 46 (II) 76 (II) 18 (I) 56 (II) 34 (II) 76 (I)

a Reaction conditions: CH Cl , reflux, 12 h (entries 1-5); 2 2 CH2Cl2, reflux, 36 h (entries 6-10); ClCH2CH2Cl, reflux, 12 h b (entries 11-13). Determined by 31P NMR. c The symbol in parentheses indicates the configuration of the major products as determined by NOE measurement and X-ray crystallography.

Figure 1. ORTEP drawing of complex 2c-I. Hydrogen atoms and counteranion PF6- are omitted for clarity.

contrast, the reaction of But analogue 1c-I produced a single diastereomer (2c-I), the configuration at the metal of which is opposite of those of the major isomers from complexes 2a and 2b (entry 5). The structure of complex 2c-I was unequivocally determined by X-ray crystallography (Figure 1). As the reactivities of trimethylphosphine (1d-f) and triphenylphosphite (1g-i) complexes toward tert-butyl isocyanide were lower than those of triphenyphosphine analogues 1a-c, slightly severe conditions were required for improving the yield of isocyanide complexes. Thus, complexes 1d-f were reacted with tert-butyl isocyanide for 36 h, and the reactions of complexes 1g-i were performed in refluxing 1,2-dichloroethane. The reactions of 1d-I and/or 1d-II gave a mixture of isocyanide complexes 2d-I and 2d-II in good yields, and the selectivity of 2d-I and 2d-II was also independent of the stereochemistry of the substrate (entries 6-8). Complex 2e-II was obtained as a major isomer from the reaction of 1e-I (entry 9). Although the structure of the major isomer produced in the reaction of 1f-I was

Ruthenium-Isocyanide and -Acetylide Complexes

Organometallics, Vol. 24, No. 11, 2005 2749

Scheme 2

Figure 2. ORTEP drawing of complex 4a-I. Hydrogen atoms and solvate are omitted for clarity. Table 2. Reactions of 1-I and/or 1-II with Phenylacetylenea

different from those of 1d-I and 1e-I, as observed in the triphenyphosphine analogue, the diastereoselectivity of 2f-I was much lower than that of 2c-I (entry 10). A similar tendency was observed for the diastereoselectivity in the reactions of 1g-i (entries 11-13). Recrystallization of a mixture of 2a-I and 2a-II in a 28:72 ratio from dichloromethane-diethyl ether led to a change of the diastereomeric ratio to 83:17. It is noteworthy that the diastereomeric ratio was not changed by refluxing the mixture of 2a-I and 2a-II (83:17) in dichloromethane for 12 h. Although similar treatments were performed in the presence of isocyanide and acetonitrile, no epimerization was detected. These results contrast the fact that epimerization between two diastereomers was observed in analogous iodo complexes.5 Next, we examined the diastereoselectivity in the formation of the ruthenium-vinylidene complex, which is known to be an important intermediate in some catalytic reactions.8 Treatment of 1a-I with 10 equiv of phenylacetylene in dichloromethane under reflux gave vinylidene complex 3a in a diastereometically pure form (Scheme 2).9 However, the isolation of vinylidene complex 3a was unsuccessful. Thus, vinylidene complex 3a was converted into acetylide complex 4a by treatment with alumina.10 Complex 4a was formed as a single diastereomer, the structure of which was determined by X-ray analysis. As shown in Figure 2, complex 4a has the configuration SCpSRu/RCpRRu (4a-I). The reaction of a mixture of the two diastereomers 1a-I and 1a-II (8) For recent reviews, see: (a) Bruneau, C.; Dixneuf, P. H. Acc. Chem. Res. 1999, 32, 311. (b) Katayama, H.; Ozawa, F. Coord. Chem. Rev. 2004, 248, 1703. (9) (a) Consiglio, G.; Morandini, F.; Ciani, G. F.; Sironi, A. Organometallics 1986, 5, 1976. (b) Consiglio, G.; Morandini, F. Inorg. Chim. Acta 1987, 127, 79. (c) Slugovc, C.; Simanko, W.; Mereiter, K.; Schmid, R.; Kirchner, K.; Xiao, L.; Weissensteiner, W. Organometallics 1999, 18, 3865. (10) Hodge, A. J.; Ingham, S. L.; Kakkar, A. K.; Khan, M. S.; Lewis, J.; Long, N. J.; Parker, D. G.; Raithby, P. R. J. Organomet. Chem. 1995, 488, 205.

entry

substrate

product

yield/%

de/%b,c

1 2 3 4 5e 6 7 8 9 10

1a-I 1a-I + 1a-II (75:25) 1b-I 1b-I + 1b-II (90:10) 1c-I 1d-I 1d-I + 1d-II (50:50) 1d-II 1e-I 1f-I

4a 4a 4b 4b 4c 4d 4d 4d 4e 4f

89 85 90 92 91 87 94 96 82 92

>99 (I) >99 (I) >99d >99d >99 (I) 74 (II) 70 (II) 74 (II) 70 (II) 40 (I)

a

Reaction conditions: CH2Cl2, reflux, 48 h. b Determined by NMR. c The symbol in parentheses indicates the configuration of the major products as determined by NOE measurement and X-ray crystallography. d The configuration of the major products could not be determined. e Reaction conditions: CH2Cl2, room temperature, 48 h.

31P

(75:25) gave complex 3a as a single diastereomer, which was also isolated after conversion into acetylide complex 4a-I in 85% yield, suggesting that the stereochemistry of the starting material does not influence the diastereoselectivity of the product. This result is similar to that observed in isocyanide complexes 2 (see above), but is in sharp contrast to the known chemistry of analogous ruthenium complexes with a chiral phosphine ligand, which stereospecifically produce the vinylidene complexes with retention of configuration at the ruthenium atom.9 Table 2 gives the representative results of the reactions of 1-I and/or 1-II with phenylacetylene followed by treatment with alumina to give acetylide complexes 4-I and/or 4-II. We also examined the diastereoselectivity of vinylidene complex 3 by measuring 31P NMR spectra of the reaction mixture before treatment with alumina and found the diastereoselectivity of 4 was essentially the same as that of 3. It seems to be reasonable that the transformation of 3 into 4 proceeds with retention of configuration at the ruthenium atom, because no bond cleavage around the ruthenium atom is required in this step. Although the reaction of complex 1b produced a single diastereomer (4b), the stereochemistry of 4b could not be determined (entries 3 and 4). Upon treatment of complex 1c-I, a single diastereomer

2750

Organometallics, Vol. 24, No. 11, 2005

Matsushima et al. Scheme 3

Scheme 4

Figure 3. ORTEP drawing of complex 4c-I. Hydrogen atoms are omitted for clarity.

(4c-I) was also obtained (entry 5), and the stereochemistry was confirmed by X-ray analysis, as shown in Figure 3. Although complexes 1d-f were also reacted with phenylacetylene to give analogous acetylide complexes (4d-f) as mixtures of two diastereomers, the configuration at the ruthenium atom of the major product was opposite that of triphenylphosphine analogues (entries 6-10). Similar reactions of complexes 1g-i did not produce any acetylide complexes at all, and the starting materials were recovered quantitatively. When the reaction was performed in refluxing 1,2-dichloroethane, the product was a complex mixture from which no acetylide complex could be isolated. A diastereomerically pure sample of 4d-II was obtained by recrystallization of a mixture of 4d-I and 4d-II (13:87) from dichloromethane-hexane. When complex 4d-II was refluxed in dichloromethane for a long time, no epimerization took place, as observed for 2a-I and 2a-II.11 Discussion Because no epimerization between the two diastereomers 2a-I and 2a-II was detected, isocyanide complex 2 was not in equilibrium depending on the thermodynamic stability of the diastereomers. Thus, the diastereoselectivity of isocyanide complex 2 is not under thermodynamic control but under kinetic control. Although it may seem unique that the diastereoselectivity was not affected by the stereochemistry of the starting material, there are two possible interpretations for this. One is that the epimerization between 1-I and 1-II is much faster than the ligand exchange reaction with isocyanide, and the other is the formation of an intermediate that has no metal-centered chirality. Although slow epimerization between 1-I and 1-II was observed for 1a and 1b, no epimerization was detected for 1d.4b The results of the reactions using 1d-I and/or 1d-II clearly suggest that the latter interpretation is plausible. Thus, a coordinatively unsaturated complex (5) is generated by the dissociation of acetonitrile, and tertbutyl isocyanide coordinates to 5, giving 2 (Scheme 3).9c (11) Carmona, D.; Vega, C.; Lahoz, F. J.; Atencio, R.; Oro, L. A.; Lamata, M. P.; Viguri, F.; Jose´, E. S. Organometallics 2000, 19, 2273.

The results in Table 1 suggest that the diastereoselectivity of 2 depended on the substituent at the 4-position of the Cp′ ring and the phosphorus ligand. The bulky substituent, But, on the Cp′ ring and the bulky phosphorus ligands, such as PPh3 and P(OPh)3, led to the increase in the selectivity of 2-I, whereas the combination of the small substituent, Me and Ph, and phosphorus ligand, PMe3, selectively produced 2-II. These steric effects may be reasonably explained by using intermediate 5, which has a two-legged piano stool structure (Scheme 4). The approach of isocyanide to the ruthenium atom of 5 from side A gives 2-I, whereas the coordination from side B produces 2-II. As one of the two phenyl rings on the tethered phosphine ligand protrudes equatorially, the approach of isocyanide from side A is affected by steric hindrance. Consequently, the approach from side B is faster than that from A, resulting in the selective formation of 2-II. This explanation is applicable to the results obtained from the combination of a small substituent and a small phosphorus ligand. Although the But group on the Cp′ ring would provide an adverse effect on the attack from side A, the selectivity of 2-I was higher than those in the reactions of 1, having a Me or a Ph group on the Cp′ ring. In particular, 2c-I was obtained as a sole product in the reaction of 1c-I, which had a bulky phosphine ligand. This inverse selectivity is likely caused by the structural change of the intermediate (5′) due to the steric repulsion between the But group on the Cp′ ring and the phosphorus ligand. Thus, the phosphorus ligand is located apart from the But group and effectively blocks the attack from side B. This explanation is consistent with the fact that the selectivity of 2-I increases with increasing bulkiness of the phosphorus ligand. The diastereoselectivity of acetylide complex 4 must be determined in the step for forming vinylidene complex 3, because the transformation of 3 into 4 proceeded with maintenance of the diastereomeric ratio and no epimerization took place in 4. Although we conducted several experiments to determine whether the diastereoselectivity of 3 was under kinetic or thermodynamic control, no conclusive results were obtained. The results in Table 2 show the steric effect of the substituent on the Cp′ ring and the phosphorus ligand on the diastereoselectivity of 4, similar to that for 2. However, if the selectivity were to depend on the thermodynamic stability of the resulting vinylidene complexes 3, a more

Ruthenium-Isocyanide and -Acetylide Complexes Scheme 5

sterically demanding substituent and phosphorus ligand would produce 3-I because the steric repulsion between the substituent on the Cp′ ring and the phosphorus ligand is important.5 The results in Table 2 are also in agreement with this explanation. Therefore, the factor contributing to the stereochemistry of 4 is still unclear. Although the stereochemistry of 4b could not be determined, the diastereoselectivity of triphenylphosphine complexes 4a-c was very high compared to that of trimethylphosphine derivatives 4d-e. This phenomenon is likely due to the steric factor of the phosphine ligand, because the selectivity of 4-I in 4f is much higher than those in 4d and 4e. It is well known that the η2-alkyne complex is generated as an intermediate in the formation of the vinylidene complex. As the coordination sphere of η2-alkyne intermediate 6 is fairly crowded, the steric effect of the phosphine ligand would be enhanced (Scheme 5). Summary We have described the diastereoselective formation of ruthenium-isocyanide and -vinylidene complexes, the latter of which were isolated as acetylide complexes. The planar-chiral cyclopentadienyl-phosphine ligand showed high ability to control the stereochemistry at the metal center in the resulting ruthenium complexes, whereas the metal-centered chirality of the starting complexes had no influence on the diastereoselectivity. A significant steric effect of the substituent on the Cp′ ring and the phosphorus ligand was observed in both reactions. Because the complexes prepared here have potential applications in other reactions, the results should be useful for developing the stereoselective transformation of isocyanides and acetylenes. Further studies that focus on the application to asymmetric catalysis are in progress. Experimental Section All reactions were carried out under an argon atmosphere, but the workup was performed in air. 1H, 13C, and 31P NMR spectra were measured on JEOL JNM-EX270, -LA400, and -LA600 spectrometers using acetone-d6 for isocyanide complexes and CDCl3 for acetylide complexes as a solvent. Chemical shifts are given in ppm based on SiMe4 as an internal standard for 1H and 13C NMR, and 85% H3PO4 as an external standard for 31P NMR. IR and mass spectra were taken on a Perkin-Elmer system 2000 FT-IR and JEOL JMS-600H instrument, respectively. Elemental analyses were performed by the Material Analysis Center, ISIR, Osaka University.

Organometallics, Vol. 24, No. 11, 2005 2751 Ruthenium complexes 1a-i were prepared by the method previously reported.4b,12 All other chemicals available commercially were used without further purification. Reaction of [(η5:η1-2-Me-4-R-C5H2CO2CH2CH2PPh2)Ru(PPh3)(NCMe)][PF6] (R ) Me, Ph, But) 1a-c with tertButyl Isocyanide. To a solution of ruthenium complex 1 (0.10 mmol) in dichloromethane (5 mL) was added tert-butyl isocyanide (80 mg, 1.0 mmol), and the reaction mixture was refluxed for 12 h. After removal of the solvent under reduced pressure, the residual oil was washed with diethyl ether several times. The crude product was purified by silica gel column chromatography using a mixture of dichloromethaneacetone (v/v ) 5/1) as eluent to give red solid. Yield and diastereoselectivity of the product 2 are given in Table 1. [(η5:η1-2,4-Me2-C5H2CO2CH2CH2PPh2)Ru(PPh3)(CNBut)][PF6] (2a-I and 2a-II). IR (cm-1, KBr): 2143 (νCtN), 1722 (νCdO). FAB-MS: m/z 796 (M - PF6-). Anal. Calcd for C45H46F6NO2P3Ru: C, 57.45; H, 4.93; N, 1.49. Found: C, 57.37; H, 4.98; N, 1.69. Major product (2a-I): 1H NMR (400 MHz): δ 7.88-7.33 (m, 17H, Ph), 6.95-6.90 (m, 6H, Ph), 6.63-6.58 (m, 2H, Ph), 5.85 (s, 1H, Cp′), 4.94 (ddd, 1H, J ) 23.7, 11.2, 7.1 Hz, OCH2), 4.49 (s, 1H, Cp′), 3.70-3.64 (m, 1H, OCH2), 3.11-3.02 (m, 1H, PCH2), 2.93-2.81 (m, 1H, PCH2), 2.52 (s, 3H, Cp′Me), 1.68 (s, 3H, Cp′Me), 1.45 (s, 9H, CMe3). 13C NMR (151 MHz): δ 165.6 (CdO), 141.6-129.1 (Ph), 117.8 (Cp′), 104.1 (Cp′), 94.3 (Cp′), 89.3 (Cp′), 83.6 (d, J ) 7 Hz, Cp′), 66.0 (CMe3), 59.7 (OCH2), 30.6 (CMe3), 25.3 (d, J ) 34 Hz, PCH2), 13.3 (Cp′Me), 12.2 (Cp′Me); the signal assignable to the isocyano-carbon could not be detected. 31P NMR (162 MHz): δ 47.7 (d, J ) 27 Hz), 36.4 (d, J ) 27 Hz). Minor product (2a-II): 1H NMR (400 MHz): δ 7.88-7.33 (m, 23H, Ph), 6.246.19 (m, 2H, Ph), 5.17-5.07 (m, 1H, OCH2), 4.84 (s, 1H, Cp′), 4.78 (s, 1H, Cp′), 4.09-4.02 (m, 1H, OCH2), 2.93-2.81 (m, 1H, PCH2), 2.72-2.63 (m, 1H, PCH2), 2.41 (s, 3H, Cp′Me), 1.35 (d, 3H, J ) 2.9 Hz, Cp′Me), 1.23 (s, 9H, CMe3). 31P NMR (162 MHz): δ 46.1 (d, J ) 26 Hz), 38.0 (d, J ) 26 Hz). [(η 5 :η 1 -2-Me-4-Ph-C 5 H 2 CO 2 CH 2 CH 2 PPh 2 )Ru(PPh 3 )(CNBut)][PF6] (2b-I and 2b-II). IR (cm-1, KBr): 2137 (νCtN), 1715 (νCdO). FAB-MS: m/z 858 (M - PF6-). Anal. Calcd for C50H48F6NO2P3Ru: C, 59.88; H, 4.82; N, 1.40. Found: C, 60.05; H, 4.92; N, 1.40. Major product (2b-I): 1H NMR (270 MHz): δ 7.76-7.22 (m, 20H, Ph), 7.01 (d, 2H, J ) 7.1 Hz, Ph), 6.60-6.40 (m, 8H, Ph), 6.15 (s, 1H, Cp′), 5.185.02 (m, 1H, OCH2), 5.10 (s, 1H, Cp′), 4.08-3.98 (m, 1H, OCH2), 3.17-3.08 (m, 2H, PCH2), 2.65 (s, 3H, Cp′Me), 1.43 (s, 9H, CMe3). 13C NMR (151 MHz): δ 165.5 (CdO), 141.5127.3 (Ph), 113.2 (Cp′), 104.7 (Cp′), 92.6 (d, J ) 10 Hz, Cp′), 91.5 (Cp′), 78.0 (d, J ) 7 Hz, Cp′), 66.0 (CMe3), 59.7 (OCH2), 30.6 (CMe3), 23.2 (d, J ) 35 Hz, PCH2), 13.5 (Cp′Me); the signal assignable to the isocyano-carbon could not be detected. 31P NMR (162 MHz): δ 42.9 (d, J ) 26 Hz), 35.1 (d, J ) 26 Hz). Minor product (2b-II): 1H NMR (270 MHz): δ 7.99 (d, 2H, J ) 7.4 Hz, Ph), 7.90 (br, 2H, Ph), 7.70-7.11 (m, 26H, Ph), 5.44 (s, 1H, Cp′), 5.37 (s, 1H, Cp′), 5.21-5.14 (m, 1H, OCH2), 4.11-4.07 (m, 1H, OCH2), 2.95-2.88 (m, 1H, PCH2), 2.77-2.73 (m, 1H, PCH2), 0.72 (s, 9H, CMe3). 31P NMR (162 MHz): δ 46.9 (d, J ) 25 Hz), 38.8 (d, J ) 25 Hz). [(η 5 :η 1 -2-Me-4-Bu t -C 5 H 2 CO 2 CH 2 CH 2 PPh 2 )Ru(PPh 3 )(CNBut)][PF6] (2c-II). 1H NMR (400 MHz): δ 8.02-7.98 (m, 2H, Ph), 7.67-7.11 (m, 21H, Ph), 6.23 (t, 2H, J ) 8.5 Hz, Ph), 5.19-5.09 (m, 1H, OCH2), 5.01-4.99 (m, 1H, Cp′), 4.514.49 (m, 1H, Cp′), 3.90-3.83 (m, 1H, OCH2), 3.18-3.09 (m, 1H, PCH2), 2.68 (dt, 1H, J ) 15.4, 4.6 Hz, PCH2), 1.59 (s, 9H, CMe3), 1.59 (d, 3H, J ) 2.7 Hz, Cp′Me), 1.14 (s, 9H, CMe3). 13C NMR (151 MHz): δ 166.8 (CdO), 147.8-129.0 (Ph), 114.1 (Cp′), 91.5 (Cp′), 87.5 (d, J ) 10 Hz, Cp′), 74.6 (d, J ) 7 Hz, Cp′), 66.0 (Cp′), 60.4 (CMe3), 60.3 (OCH2), 33.3 (CMe3), (12) Dodo, N.; Matsushima, Y.; Uno, M.; Onitsuka, K.; Takahashi, S. J. Chem. Soc., Dalton Trans. 2000, 35.

2752

Organometallics, Vol. 24, No. 11, 2005

31.2 (CMe3), 20.4 (d, J ) 30 Hz, PCH2), 12.2 (Cp′Me); the signal assignable to the isocyano carbon could not be detected. 31 P NMR (162 MHz): δ 48.0 (d, J ) 25 Hz), 33.2 (d, J ) 25 Hz). IR (cm-1, KBr): 2127 (νCtN), 1713 (νCdO). FAB-MS: m/z 838 (M - PF6-). Anal. Calcd for C48H52F6NO2P3Ru: C, 58.65; H, 5.33; N, 1.42. Found: C, 58.75; H, 5.36; N, 1.66. Reaction of [(η5:η1-2-Me-4-R-C5H2CO2CH2CH2PPh2)Ru(PMe3)(NCMe)][PF6] (R ) Me, Ph, But) (1d-f) with tertButyl Isocyanide. This reaction was performed in a manner similar to that for 1a-c with a longer reaction time (36 h). Yield and diastereoselectivity of the products are given in Table 1. [(η5:η1-2,4-Me2-C5H2CO2CH2CH2PPh2)Ru(PMe3)(CNBut)][PF6] (2d-I and 2d-II). IR (cm-1, KBr): 2126 (νCtN), 1705 (νCdO). FAB-MS: m/z 610 (M - PF6-). Anal. Calcd for C30H40F6NO2P3Ru: C, 47.75; H, 5.34; N, 1.86. Found: C, 47.90; H, 5.24; N, 1.88. Major product (2d-I): 1H NMR (600 MHz): δ 7.78-7.63 (m, 4H, Ph), 7.46 (br, 4H, Ph), 7.13 (br, 2H, Ph), 5.22 (s, 1H, Cp′), 5.11 (s, 1H, Cp′), 5.06-5.00 (m, 1H, OCH2), 3.96-3.92 (m, 1H, OCH2), 2.94-2.80 (m, 2H, PCH2), 2.42 (s, 3H, Cp′Me), 2.31 (d, 3H, J ) 2.6 Hz, Cp′Me), 1.65 (s, 9H, CMe3), 1.14 (d, 9H, J ) 9.9 Hz, PMe3). 13C NMR (151 MHz): δ 164.9 (CdO), 140.7-129.8 (Ph), 115.5 (Cp′), 104.0 (Cp′), 93.3 (Cp′), 85.6 (d, J ) 9 Hz, Cp′), 82.3 (d, J ) 6 Hz, Cp′), 59.4 (CMe3), 59.3 (OCH2), 30.8 (CMe3), 23.0 (d, J ) 35 Hz, PCH2), 20.0 (d, J ) 34 Hz, PMe3), 13.7 (Cp′Me), 13.4 (Cp′Me); the signal assignable to the isocyano carbon could not be detected. 31P NMR (162 MHz): δ 45.3 (d, J ) 32 Hz), 10.2 (d, J ) 32 Hz). Minor product (2d-II): 1H NMR (600 MHz): δ 7.887.85 (m, 2H, Ph), 7.78-7.63 (m, 6H, Ph), 7.37 (br, 2H, Ph), 5.18 (s, 1H, Cp′), 5.06-5.00 (m, 1H, OCH2), 4.95 (s, 1H, Cp′), 3.96-3.92 (m, 1H, OCH2), 3.21-3.15 (m, 1H, PCH2), 2.942.80 (m, 1H, PCH2), 2.39 (d, 3H, J ) 3.9 Hz, Cp′Me), 2.34 (s, 3H, Cp′Me), 1.54 (s, 9H, CMe3), 1.12 (d, 9H, J ) 9.9 Hz, PMe3). 31P NMR (162 MHz): δ 43.1 (d, J ) 31 Hz), 10.8 (d, J ) 31 Hz). [(η 5 :η 1 -2-Me-4-Ph-C 5 H 2 CO 2 CH 2 CH 2 PPh 2 )Ru(PMe 3 )(CNBut)][PF6] (2e-I and 2e-II). IR (cm-1, KBr): 2128 (νCtN), 1715 (νCdO). FAB MS: m/z 672 (M - PF6-). Anal. Calcd for C35H42F6NO2P3Ru: C, 51.47; H, 5.18; N, 1.72. Found: C, 51.47; H, 5.06; N, 1.64. Major product (2e-I): 1H NMR (400 MHz): δ 7.92 (d, 2H, J ) 7.3 Hz, Ph), 7.77-7.73 (m, 5H, Ph), 7.64 (t, 2H, J ) 7.3 Hz, Ph), 7.55 (t, 1H, J ) 7.3 Hz, Ph), 7.44-7.43 (m, 3H, Ph), 7.11-7.07 (m, 2H, Ph), 6.00 (s, 1H, Cp′), 5.79 (s, 1H, Cp′), 5.15-5.05 (m, 1H, OCH2), 4.12-4.08 (m, 1H, OCH2), 3.00-2.95 (m, 2H, PCH2), 2.55 (s, 3H, Cp′Me), 1.67 (s, 9H, CMe3), 0.82 (d, 9H, J ) 10.3 Hz, PMe3). 13C NMR (151 MHz): δ 165.0 (CdO), 140.4-127.2 (Ph), 115.9 (Cp′), 106.2 (Cp′), 89.4 (Cp′), 84.0 (d, J ) 6 Hz, Cp′), 80.8 (Cp′), 59.7 (CMe3), 59.6 (d, J ) 5 Hz, OCH2), 30.2 (CMe3), 22.8 (d, J ) 35 Hz, PCH2), 18.8 (d, J ) 34 Hz, PMe3), 14.0 (Cp′Me); the signal assignable to the isocyano carbon could not be detected. 31P NMR (162 MHz): δ 44.6 (d, J ) 32 Hz), 10.5 (d, J ) 32 Hz). Minor product (2e-II): 31P NMR (162 MHz): δ 44.0 (d, J ) 30 Hz), 12.1 (d, J ) 30 Hz). [(η 5 :η 1 -2-Me-4-Bu t -C 5 H 2 CO 2 CH 2 CH 2 PPh 2 )Ru(PPh 3 )(CNBut)][PF6] (2f-I and 2f-II). IR (cm-1, KBr): 2112 (νCtN), 1715 (νCdO). FAB-MS: m/z 652 (M - PF6-). Anal. Calcd for C33H46F6NO2P3Ru: C, 49.75; H, 5.82; N, 1.76. Found: C, 49.55; H, 5.74; N, 1.74. Major product (2f-II): 1H NMR (400 MHz): δ 8.00-7.95 (m, 2H, Ph), 7.63-7.47 (m, 6H, Ph), 7.37 (br, 2H, Ph), 5.19 (s, 1H, Cp′), 5.17 (s, 1H, Cp′), 5.07-4.97 (m, 1H, OCH2), 3.76-3.69 (m, 1H, OCH2), 3.08-2.81 (m, 2H, PCH2), 2.40 (s, 3H, Cp′Me), 1.51 (s, 9H, CMe3), 1.50 (s, 9H, CMe3), 1.14 (d, 9H, J ) 10.3 Hz, PMe3). 13C NMR (151 MHz): δ 166.5 (CdO), 140.9-129.2 (Ph), 111.7 (Cp′), 92.4 (Cp′), 89.9 (Cp′), 80.3 (d, J ) 7 Hz, Cp′), 77.1 (d, J ) 6 Hz, Cp′), 60.3 (CMe3), 59.9 (OCH2), 33.8 (CMe3), 32.1 (CMe3), 31.3 (CMe3), 22.4 (d, J ) 34 Hz, PCH2), 18.8 (d, J ) 32 Hz, PMe3), 13.2 (Cp′Me); the signal assignable to the isocyano carbon could not be

Matsushima et al. detected. 31P NMR (162 MHz): δ 41.7 (d, J ) 31 Hz), 8.3 (d, J ) 31 Hz). Minor product (2f-I): 1H NMR (400 MHz): δ 7.91-7.86 (m, 2H, Ph), 7.70-7.68 (m, 1H, Ph), 7.63-7.47 (m, 5H, Ph), 7.19 (br, 2H, Ph), 5.38 (s, 1H, Cp′), 5.27 (s, 1H, Cp′), 5.07-4.97 (m, 1H, OCH2), 4.05 (br, 1H, OCH2), 3.082.81 (m, 2H, PCH2), 2.41 (s, 3H, Cp′Me), 1.66 (s, 9H, CMe3), 1.45 (s, 9H, CMe3), 1.21 (d, 9H, J ) 9.8 Hz, PMe3). 31P NMR (162 MHz): δ 40.3 (d, J ) 30 Hz), 10.4 (d, J ) 30 Hz). Reaction of [(η5:η1-2-Me-4-R-C5H2CO2CH2CH2PPh2)Ru{P(OPh)3}(NCMe)][PF6] (R ) Me, Ph, But) (1g-i) with tert-Butyl Isocyanide. Ruthenium complexes 1g-i (0.1 mmol) were treated with tert-butyl isocyanide (0.08 g, 1.0 mmol) in refluxing 1,2-dichloroethane (5 mL) for 12 h. Purification similar to that for 1a-c gave the products in yields with diastereoselectivity given in Table 1, respectively. [(η 5 :η 1 -2,4-Me2 -C 5 H 2 CO 2 CH 2 CH 2 PPh 2 )Ru{P(OPh)3 }(CNBut)][PF6] (2g-I and 2g-II). IR (cm-1, KBr): 2161 (νCtN), 1722 (νCdO). FAB-MS: m/z 844 (M - PF6-). Anal. Calcd for C45H46F6NO5P3Ru: C, 54.66; H, 4.69; N, 1.42. Found: C, 54.44; H, 4.66; N, 1.67. Major product (2g-I): 1H NMR (600 MHz): δ 7.98-7.95 (m, 2H, Ph), 7.68-7.67 (m, 3H, Ph), 7.55-7.52 (m, 2H, Ph), 7.48-7.46 (m, 1H, Ph), 7.41-7.38 (m, 2H, Ph), 7.24-7.21 (m, 6H, Ph), 7.12 (t, 3H, J ) 7.5 Hz, Ph), 6.76 (d, 6H, J ) 8.6 Hz, Ph), 5.19 (s, 1H, Cp′), 5.14-5.08 (m, 1H, OCH2), 4.14-4.09 (m, 1H, OCH2), 4.07 (s, 1H, Cp′), 3.09-3.02 (m, 2H, PCH2), 2.43 (d, 3H, J ) 3.8 Hz, Cp′Me), 2.23 (s, 3H, Cp′Me), 1.71 (s, 9H, CMe3). 13C NMR (151 MHz): δ 164.6 (CdO), 152.6 (CtN), 140.1-121.2 (Ph), 116.5 (Cp′), 106.0 (Cp′), 94.2 (Cp′), 86.0 (d, J ) 7 Hz, Cp′), 83.6 (d, J ) 14 Hz, Cp′), 60.2 (CMe3), 59.8 (d, J ) 5 Hz, OCH2), 30.9 (CMe3), 23.2 (d, J ) 35 Hz, PCH2), 13.4 (Cp′Me), 13.3 (Cp′Me). 31P NMR (162 MHz): δ 133.2 (d, J ) 52 Hz), 42.9 (d, J ) 52 Hz). Minor product (2g-II): 1H NMR (600 MHz): δ 2.41 (s, 3H, Cp′Me), 2.26 (d, 3H, J ) 3.4 Hz, Cp′Me), 1.64 (s, 9H, CMe3); the signals assignable to the other protons could not be identified due to overlap with those of the major product. 31P NMR (162 MHz): δ 132.2 (d, J ) 55 Hz); the other signal could not be detected due to overlap with that of the major product. [(η5:η1-2-Me-4-Ph-C5H2CO2CH2CH2PPh2)Ru{P(OPh)3}(CNBut)][PF6] (2h-I and 2h-II). 13C NMR (151 MHz): δ 164.9 (CdO), 164.6 (CdO), 152.4 (d, J ) 15 Hz, CtN), 152.3 (d, J ) 14 Hz, CtN), 139.9-120.8 (Ph), 116.7 (Cp′), 113.2 (d, J ) 5 Hz, Cp′), 108.0 (Cp′), 91.8 (Cp′), 83.5 (d, J ) 13 Hz, Cp′), 82.7 (d, J ) 13 Hz, Cp′), 82.3 (d, J ) 7 Hz, Cp′), 78.7 (Cp′), 60.5 (CMe3), 60.0 (d, J ) 5 Hz, OCH2), 60.3 (d, J ) 5 Hz, OCH2), 30.9 (CMe3), 30.4 (CMe3), 22.9 (d, J ) 35 Hz, PCH2), 20.2 (d, J ) 34 Hz, PCH2), 13.5 (Cp′Me), 13.3 (Cp′Me). IR (cm-1, KBr): 2151 (νCtN), 1727 (νCdO). FAB-MS: m/z 906 (M - PF6-). Anal. Calcd for C50H48F6NO5P3Ru: C, 57.15; H, 4.60; N, 1.33. Found: C, 57.21; H, 4.82; N, 1.34. Major product (2h-I): 1H NMR (400 MHz): δ 7.90-7.85 (m, 2H, Ph), 7.73-7.29 (m, 13H, Ph), 7.15 (t, 6H, J ) 7.3 Hz, OPh), 7.07 (t, 3H, J ) 7.3 Hz, OPh), 6.41 (d, 6H, J ) 8.5 Hz, OPh), 5.86 (d, 1H, J ) 1.7 Hz, Cp′), 5.18 (d, 1H, J ) 1.7 Hz, Cp′), 5.275.15 (m, 1H, OCH2), 4.34-4.26 (m, 1H, OCH2), 3.15-3.03 (m, 2H, PCH2), 2.44 (s, 3H, Cp′Me), 1.75 (s, 9H, CMe3). 31P NMR (162 MHz): δ 128.3 (d, J ) 51 Hz), 42.1 (d, J ) 51 Hz). Minor product (2h-II): 1H NMR (400 MHz): δ 7.847.78 (m, 2H, Ph), 7.73-7.29 (m, 19H, Ph, OPh), 7.21 (t, 3H, J ) 7.3 Hz, OPh), 6.85 (d, 6H, J ) 8.5 Hz, OPh), 5.81 (m, 1H, Cp′), 5.27-5.15 (m, 1H, OCH2), 4.67 (s, 1H, Cp′), 4.18-4.10 (m, 1H, OCH2), 3.35-3.25 (m, 2H, PCH2), 2.56 (d, 3H, J ) 8.5 Hz, Cp′Me), 1.16 (s, 9H, CMe3). 31P NMR (162 MHz): δ 132.3 (d, J ) 53 Hz), 44.3 (d, J ) 53 Hz). [(η5:η1-2-Me-4-But-C5H2CO2CH2CH2PPh2)Ru{P(OPh)3}(CNBut)][PF6] (2i-I and 2i-II). IR (cm-1, KBr): 2144 (νCtN), 1732 (νCdO). FAB-MS: m/z 886 (M - PF6-). Anal. Calcd for C48H52F6NO5P3Ru: C, 55.92; H, 5.08; N, 1.36. Found: C, 55.72; H, 5.00; N, 1.16. Major product (2i-II): 1H NMR (400 MHz): δ 8.02-7.99 (m, 2H, Ph), 7.69-7.48 (m, 8H, Ph), 7.25 (t, 6H,

Ruthenium-Isocyanide and -Acetylide Complexes J ) 8.6 Hz, OPh), 7.15 (t, 3H, J ) 7.5 Hz, OPh), 6.82 (d, 6H, J ) 8.6 Hz, OPh), 5.37 (m, 1H, Cp′), 5.17-5.10 (m, 1H, OCH2), 4.20 (s, 1H, Cp′), 3.94-3.89 (m, 1H, OCH2), 3.34-3.29 (m, 1H, PCH2), 3.19 (dt, 1H, J ) 15.4, 4.2 Hz, PCH2), 2.50 (d, 3H, J ) 5.9 Hz, Cp′Me), 1.57 (s, 9H, CMe3), 1.37 (s, 9H, CMe3). 13 C NMR (151 MHz): δ 166.0 (CdO), 152.7 (d, J ) 14 Hz, CN), 140.6-121.2 (Ph), 114.2 (Cp′), 92.6 (Cp′), 79.6 (d, J ) 4 Hz, Cp′), 79.4 (d, J ) 7 Hz, Cp′), 60.7 (d, J ) 5 Hz, OCH2), 60.6 (CMe3), 32.9 (CMe3), 31.1 (CMe3), 30.08 (CMe3), 21.1 (d, J ) 35 Hz, PCH2), 13.4 (Cp′Me). 31P NMR (162 MHz): δ 131.7 (d, J ) 55 Hz), 40.0 (d, J ) 55 Hz). Minor product (2i-I): 1 H NMR (600 MHz): δ 6.75 (d, 6H, J ) 7.8 Hz, OPh), 2.25 (s, 3H, Cp′Me), 1.77 (s, 9H, CMe3), 1.43 (s, 9H, CMe3); the signals assignable to the other protons could not be identified due to overlap with those of the major product. 31P NMR (162 MHz): δ 132.5 (d, J ) 50 Hz), 38.4 (d, J ) 50 Hz). Reaction of Ruthenium Complexes 1a-f with Phenylacetylene. To a dichloromethane solution (5 mL) of ruthenium complex 1 (0.1 mmol) was added phenylacetylene (0.11 g, 1.0 mmol), and the reaction mixture was refluxed for 48 h. The solvent was evaporated, and the residual oil was washed with diethyl ether several times. The residue was dissolved in dichloromethane and placed on an alumina column. The benzene elution was collected, and the solvent was removed under reduced pressure to give a yellow solid. Yield and diastereoselectivity of the product 4 are given in Table 2. The reaction of 1c was performed at room temperature for 4 days because no vinylidene complex was formed in the reaction in refluxing dichloromethane. (η 5 :η 1 -2,4-Me 2 -C 5 H 2 CO 2 CH 2 CH 2 PPh 2 )Ru(PPh 3 )(Ct CPh) (4a-I). 1H NMR (400 MHz): δ 8.02-8.00 (m, 2H, Ph), 7.39-6.79 (m, 26H, Ph), 6.38 (t, 2H, J ) 8.3 Hz, Ph), 5.06 (ddd, 1H, J ) 19.5, 11.2, 8.3 Hz, OCH2), 4.52 (d, 1H, J ) 2.2 Hz, Cp′), 3.99 (d, 1H, J ) 2.2 Hz, Cp′), 3.89-3.84 (m, 1H, OCH2), 2.67-2.58 (m, 1H, PCH2), 2.35-2.34 (s, 3H, Cp′Me), 2.14 (dt, 1H, J ) 14.9, 5.1 Hz, PCH2), 1.39 (s, 3H, Cp′Me). 13C NMR (151 MHz): δ 166.3 (CdO), 142.6-127.4 (Ph), 127.1 (Cp′), 123.0 (Cp′), 110.8 (d, J ) 4 Hz, Cp′), 109.9 (d, J ) 20 Hz, RuCt C), 87.0 (d, J ) 11 Hz, RuCtC), 84.7 (Cp′), 81.3 (d, J ) 9 Hz, Cp′), 58.1 (d, J ) 4 Hz, OCH2), 18.5 (d, J ) 25 Hz, PCH2), 14.2 (Cp′Me), 11.8 (Cp′Me). 31P NMR (162 MHz): δ 47.5 (d, J ) 31 Hz), 35.3 (d, J ) 31 Hz). IR (cm-1, KBr): 2076 (νCtC), 1704 (νCdO). FAB-MS: m/z 814 (M+). Anal. Calcd for C48H42O2P2Ru: C, 70.84; H, 5.20. Found: C, 70.62; H, 5.27. (η5:η1-2-Me-4-Ph-C5H2CO2CH2CH2PPh2)Ru(PPh3)(Ct CPh) (4b). 1H NMR (400 MHz): δ 7.80-7.76 (m, 2H, Ph), 7.66-7.64 (m, 2H, Ph), 7.48-6.77 (m, 27H, Ph), 6.33 (t, 2H, J ) 8.5 Hz, Ph), 6.26 (d, 2H, J ) 7.3 Hz, Ph), 5.16-5.06 (m, 1H, OCH2), 5.09 (s, 1H, Cp′), 4.50 (s, 1H, Cp′), 3.96 (dt, 1H, J ) 10.3, 5.1 Hz, OCH2), 2.74-2.66 (m, 1H, CH2P), 2.20 (dt, 1H, J ) 14.6, 5.1 Hz, PCH2), 1.50 (s, 3H, Cp′Me). 13C NMR (100 MHz): δ 166.2 (CdO), 142.3-127.0 (Ph), 122.7 (Cp′), 113.1 (d, J ) 5 Hz, Cp′), 110.8 (RuCtC or Cp′), 110.1(RuCtC or Cp′), 86.2 (Cp′), 84.9 (d, J ) 13 Hz, RuCtC), 78.3 (d, J ) 8 Hz, Cp′), 58.3 (d, J ) 4 Hz, OCH2), 20.0 (d, J ) 26 Hz, PCH2), 12.1 (Cp′Me). 31P NMR (162 MHz): δ 47.8 (d, J ) 31 Hz), 36.2 (d, J ) 31 Hz). IR (cm-1, KBr): 2080 (νCtC), 1708 (νCdO). FAB-MS: m/z 876 (M+). Anal. Calcd for C53H44O2P2Ru: C, 72.59; H, 5.17. Found: C, 72.39; H, 4.88. (η5:η1-2-Me-4-But-C5H2CO2CH2CH2PPh2)Ru(PPh3)(Ct CPh) (4c-I). 1H NMR (400 MHz): δ 8.08 (t, 3H, J ) 8.4 Hz, Ph), 7.38-7.25 (m, 16H, Ph), 7.04 (t, 2H, J ) 7.5 Hz, Ph), 6.99 (t, 2H, J ) 7.5 Hz, Ph), 6.91 (t, 1H, J ) 7.1 Hz, Ph), 6.81 (d, 2H, J ) 8.4 Hz, Ph), 6.73 (t, 2H, J ) 7.5 Hz, Ph), 6.49 (t, 2H, J ) 8.4 Hz, Ph), 5.07 (ddd, 1H, J ) 19.0, 11.2, 7.9 Hz, OCH2), 4.77 (s, 1H, Cp′), 3.72 (s, 1H, Cp′), 3.68-3.64 (m, 1H, OCH2), 3.02-2.96 (m, 1H, PCH2), 2.14 (dt, 1H, J ) 14.5, 4.8 Hz, PCH2), 1.63 (s, 3H, Cp′Me), 1.59 (s, 9H, CMe3). 13C NMR (151 MHz): δ 167.8 (CdO), 142.8-127.4 (Ph), 127.2 (d, J ) 10 Hz, Cp′), 122.8 (Cp′), 114.3 (d, J ) 11 Hz, RuCtC), 105.8 (Cp′), 86.9 (d, J ) 11 Hz, RuCtC), 86.4 (Cp′), 71.7

Organometallics, Vol. 24, No. 11, 2005 2753 (d, J ) 10 Hz, Cp′), 58.8 (OCH2), 33.0 (CMe3), 30.3 (CMe3), 19.5 (d, J ) 26 Hz, PCH2), 12.3 (Cp′Me). 31P NMR (162 MHz): δ 44.4 (d, J ) 29 Hz), 21.7 (d, J ) 29 Hz). IR (cm-1, KBr): 2081 (νCtC), 1713 (νCdO). FAB-MS: m/z 856 (M+). Anal. Calcd for C51H48O2P2Ru: C, 71.56; H, 5.65. Found: C, 71.56; H, 5.65. (η 5 :η 1 -2,4-Me 2 -C 5 H 2 CO 2 CH 2 CH 2 PPh 2 )Ru(PMe 3 )(Ct CPh) (4d). IR (cm-1, KBr): 2076 (νCtC), 1697 (νCdO). FABMS: m/z 628 (M+). Anal. Calcd for C33H36O2P2Ru: C, 63.15; H, 5.78. Found: C, 63.14; H, 5.68. Major product (4d-II): 1 H NMR (600 MHz): δ 7.65 (t, 2H, J ) 8.1 Hz, Ph), 7.52-7.48 (m, 3H, Ph), 7.31 (d, 2H, J ) 8.1 Hz, Ph), 7.26-7.15 (m, 6H, Ph), 7.04 (t, 1H, J ) 7.3 Hz, Ph), 5.02-4.95 (m, 1H, OCH2), 4.77 (s, 1H, Cp′), 4.58 (s, 1H, Cp′), 3.79-3.74 (m, 1H, OCH2), 3.31-3.24 (m, 1H, PCH2), 2.34 (s, 3H, Cp′Me), 2.27-2.22 (m, 1H, PCH2), 2.22 (s, 3H, Cp′Me), 1.00 (d, 9H, J ) 9.2 Hz, PMe3). 13C NMR (151 MHz): δ 164.7 (CdO), 142.2-127.5 (Ph), 123.4 (RuCtC), 114.4 (Cp′), 105.6 (RuCtC), 94.8 (Cp′), 89.2 (Cp′), 84.5 (d, J ) 10 Hz, Cp′), 79.5 (d, J ) 9 Hz, Cp′), 58.3 (d, J ) 4 Hz, OCH2), 22.1 (d, J ) 34 Hz, PCH2), 20.6 (d, J ) 30 Hz, PMe3), 14.2 (Cp′Me), 13.6 (Cp′Me). 31P NMR (162 MHz): 45.1 (d, J ) 39 Hz), 9.7 (d, J ) 39 Hz). Minor product (4d-I): 1H NMR (600 MHz): δ 7.91-7.88 (m, 2H, Ph), 7.74-7.73 (m, 1H, Ph), 7.41-6.95 (m, 12H, Ph), 5.07 (ddd, 1H, J ) 20.1, 10.7, 7.1 Hz, OCH2), 4.61 (s, 1H, Cp′), 4.53 (s, 1H, Cp′), 3.87-3.82 (m, 1H, OCH2), 2.76-2.64 (m, 2H, PCH2), 2.35 (d, 3H, J ) 3.0 Hz, Cp′Me), 2.30 (d, 3H, J ) 1.6 Hz, Cp′Me), 0.98 (d, 9H, J ) 9.1 Hz, PMe3). 31P NMR (162 MHz): δ 38.2 (d, J ) 36 Hz), 9.5 (d, J ) 36 Hz). (η5:η1-2-Me-4-Ph-C5H2CO2CH2CH2PPh2)Ru(PMe3)(Ct CPh) (4e). IR (cm-1, KBr): 2084 (νCtC), 1704 (νCdO). FABMS: m/z 690 (M+). Anal. Calcd for C38H38O2P2Ru: C, 66.17; H, 5.55. Found: C, 66.44; H, 5.35. Major product (4e-II): 1H NMR (600 MHz): δ 7.66-6.69 (m, 20H, Ph), 5.37 (s, 1H, Cp′), 5.17 (s, 1H, Cp′), 5.10-5.03 (m, 1H, OCH2), 3.95-3.90 (m, 1H, OCH2), 3.34-3.27 (m, 1H, PCH2), 2.50 (s, 3H, Cp′Me), 2.27 (dt, 1H, J ) 14.6, 4.9 Hz, PCH2), 0.71 (d, 9H, J ) 9.6 Hz, PMe3). 13C NMR (151 MHz): 164.7 (CdO), 142.0-125.0 (Ph), 123.5 (RutCC), 114.7 (Cp′), 105.7 (RuC≡C), 97.2 (Cp′), 86.2 (Cp′), 80.8 (d, J ) 7 Hz, Cp′), 79.7 (d, J ) 11 Hz, Cp′), 58.5 (d, J ) 3 Hz, OCH2), 21.7 (d, J ) 35 Hz, PCH2), 19.5 (d, J ) 31 Hz, PMe3), 13.9 (Cp′Me). 31P NMR (162 MHz): 43.6 (d, J ) 38 Hz), 9.5 (d, J ) 38 Hz). Minor product (4e-I): 1H NMR (600 MHz): δ 7.66-6.69 (m, 20H, Ph), 5.19 (s, 1H, Cp′), 5.09 (s, 1H, Cp′), 5.10-5.03 (m, 1H, OCH2), 4.05-4.00 (m, 1H, OCH2), 2.79-2.69 (m, 2H, PCH2), 2.42 (d, 3H, J ) 2.5 Hz, Cp′Me), 1.03 (d, 9H, J ) 9.1 Hz, PMe3). 31P NMR (162 MHz): δ 39.1 (d, J ) 36 Hz), 10.6 (d, J ) 36 Hz). (η5:η1-2-Me-4-But-C5H2CO2CH2CH2PPh2)Ru(PMe3)(Ct CPh) (4f). IR (cm-1, KBr): 2080 (νCtC), 1714 (νCdO). FAB-MS: m/z 670 (M+). Anal. Calcd for C36H42O2P2Ru: C, 64.56; H, 6.32. Found: C, 65.02; H, 6.26. Major product (4f-I): 1H NMR (400 MHz): δ 8.02-7.97 (m, 2H, Ph), 7.50-6.92 (m, 13H, Ph), 5.04-4.94 (m, 1H, OCH2), 4.87 (s, 1H, Cp′), 4.53 (s, 1H, Cp′), 3.64-3.57 (m, 1H, OCH2), 2.82-2.74 (m, 1H, PCH2), 2.53 (dt, 1H, J ) 14.1, 4.4 Hz, PCH2), 2.37 (s, 3H, Cp′Me), 1.54 (s, 9H, CMe3), 0.95 (d, 9H, J ) 9.1 Hz, PMe3). 13C NMR (151 MHz): δ 167.7 (CdO), 143.0-127.3 (Ph), 123.4 (RuCt C), 115.3 (Cp′), 110.7 (RuCtC), 103.9 (Cp′), 87.2 (Cp′), 78.2 (d, J ) 10 Hz, Cp′), 73.9 (d, J ) 8 Hz, Cp′), 58.5 (d, J ) 3 Hz, OCH2), 32.5 (CMe3), 30.7 (CMe3), 22.8 (d, J ) 31 Hz, PCH2), 19.0 (d, J ) 31 Hz, PMe3), 13.7 (Cp′Me). 31P NMR (162 MHz): δ 35.9 (d, J ) 35 Hz), 9.0 (d, J ) 35 Hz). Minor product (4f-II): 1H NMR (400 MHz): δ 7.85-7.79 (m, 2H, Ph), 7.506.92 (m, 13H, Ph), 5.04-4.94 (m, 1H, OCH2), 4.87 (s, 1H, Cp′), 4.59 (s, 1H, Cp′), 3.79-3.73 (m, 1H, OCH2), 3.45-3.37 (m, 1H, PCH2), 2.39 (s, 3H, Cp′Me), 2.20 (dt, 1H, J ) 14.4, 4.6 Hz, PCH2), 1.37 (s, 9H, CMe3), 1.04 (d, 9H, J ) 9.0 Hz, PMe3). 31P NMR (162 MHz): δ 45.41 (d, J ) 36 Hz), 7.6 (d, J ) 36 Hz). X-ray Diffraction Analyses of Complexes 2c-I, 4a-I, and 4c-I. Crystals suitable for X-ray diffraction were mounted

2754

Organometallics, Vol. 24, No. 11, 2005

Matsushima et al.

Table 3. Summary of Crystallographic Data for Complexes 2c-I, 4a-I, and 4c-I 2c-I empirical formula fw cryst dimens/mm cryst syst lattice params a/Å b/Å c/Å R/deg β/deg γ/deg V/Å3 space group Z value Dcalcd/g cm-3 F(000) µ(Mo KR)/cm-1 no. reflns measd total unique no. observations no. params residuals: R1; wR2 GOF peak, hole/e Å-3

4a-I‚0.5CHCl3

4c-I‚CH2Cl2

C48H52F6NO2P3Ru 982.93 0.30 × 0.20 × 0.05 monoclinic

C48.5H42.5Cl1.5O2P2Ru 873,57 0.50 × 0.25 × 0.05 triclinic

C52H50Cl2O2P2Ru 982.93 0.25 × 0.20 × 0.15 monoclinic

13.739(6) 18.886(6) 17.871(5)

9.875(2) 21.578(2) 21.305(2)

4628(2) P21/n (# 14) 4 1.411 2024 5.05

12.584(6) 19.703(9) 9.045(7) 98.03(6) 104.57(3) 72.12(4) 2060(2) P1 h (# 2) 2 1.408 898 5.94

4441(1) P21/c (# 14) 4 1.407 1944 5.86

11 450 11 106 (Rint ) 0.160) 5661 (I > 2.0σ(I)) 602 0.054; 0.114 1.098 0.86, -1.02

9954 9518 (Rint ) 0.094) 7843 (I > 2.0σ(I)) 554 0.047; 0.082 1.186 0.92, -1.51

9100 8844 (Rint ) 0.160) 5016 (I > 2.0σ(I)) 582 0.065; 0.130 1.177 1.25, -1.05

93.53(3)

on a glass fiber with epoxy resin. All measurements were performed on a Rigaku AFC7R or AFC5R automated fourcircle diffractometer using graphite-monochromated Mo KR radiation (λ ) 0.71069 Å). A summary of crystallographic data is given in Table 3. Additional information on the collection of the data and the refinement of the structures is available as Supporting Information.

Acknowledgment. This work was supported in part by a Grant-in-Aid for Scientific Research from the

101.92(1)

Ministry of Education, Science, Sports and Culture. We thank the members of the Material Analysis Center, ISIR, Osaka University, for spectral measurements, X-ray diffraction, and microanalyses. Supporting Information Available: Details of crystallographic work (CIF file). This material is available free of charge via the Internet at http://pubs.acs.org. OM0500523